首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The reaction of superoxide anions with myeloperoxidase (donor: hydrogen-peroxide oxidoreductase, EC 1.11.1.7), which results in the formation of Compound III of myeloperoxidase, was investigated. It is shown that myeloperoxidase has a high affinity for superoxide anions because formation of Compound III was only partially inhibited by high concentrations of superoxide dismutase. Furthermore, when superoxide anions were generated in a mixture of both cytochrome c and myeloperoxidase in the absence of Cl-, only Compound III was formed and reduction of cytochrome c was not observed. In the presence of Cl-, Compound III was also formed and reduction of cytochrome c was inhibited. From the results described in this paper we conclude that Compound III is able to react with superoxide anions, probably resulting in formation of an intermediate (Compound I) which is catalytically active in the oxidation of Cl- to yield hypochlorous acid (HOCl). Because Compound III of myeloperoxidase is formed in phagocytosing neutrophils (Winterbourn, C.C., Garcia, R.C. and Segal, A.W. (1985) Biochem. J. 228, 583-592) we propose that, in vivo, myeloperoxidase also acts as a superoxide dismutase, and via formation of Compound I uses superoxide anions in the formation of HOCl.  相似文献   

2.
We investigated the effect of D-penicillamine on the ability of myeloperoxidase, purified from human leukocytes, to catalyse the oxidation of chloride ions to hypochlorite (HOCl) in the presence of H2O2. It is shown that, due to the interaction of D-penicillamine with both myeloperoxidase itself and HOCl, the chlorinating activity of myeloperoxidase in the presence of H2O2 and chloride ions is prevented. A concentration of 100 microM D-penicillamine inhibits the chlorinating activity of myeloperoxidase completely, which Is due to the stabilization of Compound II, an inactive form of the enzyme. In addition, HOCl reacts directly with D-penicillamine. Analysis of the reaction products of D-penicillamine and HOCl showed that D-penicillamine was oxidized to penicillamine disulphide and penicillamine sulphinic acid, and eventually deaminated (indicated by the release of ammonia). Lower concentrations of D-penicillamine (10 microM) inhibited myeloperoxidase less, but still acted as effective scavengers of HOCl. In very low concentrations (1 microM), D-penicillamine did not scavenge HOCl effectively, but rather stimulated the chlorinating activity of myeloperoxidase. However, when instead of D-penicillamine a comparable amount of ascorbate was added, a similar but even larger stimulation was observed. Since the concentration of free D-penicillamine in serum from rheumatoid patients treated with this drug is about 20 microM (Saetre, R. and Rabenstein, D.L. (1978) Anal. Chem. 50, 276-280), the therapeutic effect of D-penicillamine may be due to the protection of tissues against the reactive HOCl released by activated granulocytes at inflammation sites.  相似文献   

3.
Stimulated neutrophils discharge large quantities of superoxide (O2.-), which dismutates to form H2O2. In combination with Cl-, H2O2 is converted into the potent oxidant hypochlorous acid (HOCl) by the haem enzyme myeloperoxidase. We have used an H2O2 electrode to monitor H2O2 uptake by myeloperoxidase, and have shown that in the presence of Cl- this accurately represents production of HOCl. Monochlorodimedon, which is routinely used to assay production of HOCl, inhibited H2O2 uptake by 95%. This result confirms that monochlorodimedon inhibits myeloperoxidase, and that the monochlorodimedon assay grossly underestimates the activity of myeloperoxidase. With 10 microM-H2O2 and 100 mM-Cl-, myeloperoxidase had a neutral pH optimum. Increasing the H2O2 concentration to 100 microM lowered the pH optimum to pH 6.5. Above the pH optimum there was a burst of H2O2 uptake that rapidly declined due to accumulation of Compound II. High concentrations of H2O2 inhibited myeloperoxidase and promoted the formation of Compound II. These effects of H2O2 were decreased at higher concentrations of Cl-. We propose that H2O2 competes with Cl- for Compound I and reduces it to Compound II, thereby inhibiting myeloperoxidase. Above pH 6.5, O2.- generated by xanthine oxidase and acetaldehyde prevented H2O2 from inhibiting myeloperoxidase, increasing the initial rate of H2O2 uptake. O2.- allowed myeloperoxidase to function optimally with 100 microM-H2O2 at pH 7.0. This occurred because, as previously demonstrated, O2.- prevents Compound II from accumulating by reducing it to ferric myeloperoxidase. In contrast, at pH 6.0, where Compound II did not accumulate, O2.- retarded the uptake of H2O2. We propose that by generating O2.- neutrophils prevent H2O2 and other one-electron donors from inhibiting myeloperoxidase, and ensure that this enzyme functions optimally at neutral pH.  相似文献   

4.
Neutrophils kill bacteria by ingesting them into phagosomes where superoxide and cytoplasmic granule constituents, including myeloperoxidase, are released. Myeloperoxidase converts chloride and hydrogen peroxide to hypochlorous acid (HOCl), which is strongly microbicidal. However, the role of oxidants in killing and the species responsible are poorly understood and the subject of current debate. To assess what oxidative mechanisms are likely to operate in the narrow confines of the phagosome, we have used a kinetic model to examine the fate of superoxide and its interactions with myeloperoxidase. Known rate constants for reactions of myeloperoxidase have been used and substrate concentrations estimated from neutrophil morphology. In the model, superoxide is generated at several mm/s. Most react with myeloperoxidase, which is present at millimolar concentrations, and rapidly convert the enzyme to compound III. Compound III turnover by superoxide is essential to maintain enzyme activity. Superoxide stabilizes at approximately 25 microM and hydrogen peroxide in the low micromolar range. HOCl production is efficient if there is adequate chloride supply, but further knowledge on chloride concentrations and transport mechanisms is needed to assess whether this is the case. Low myeloperoxidase concentrations also limit HOCl production by allowing more hydrogen peroxide to escape from the phagosome. In the absence of myeloperoxidase, superoxide increases to >100 microM but hydrogen peroxide to only approximately 30 microM. Most of the HOCl reacts with released granule proteins before reaching the bacterium, and chloramine products may be effectors of its antimicrobial activity. Hydroxyl radicals should form only after all susceptible protein targets are consumed.  相似文献   

5.
Phagocytes generate superoxide (O2-.) and hydrogen peroxide (H2O2) and their interaction in an iron-catalyzed reaction to form hydroxyl radicals (OH.) (Haber-Weiss reaction) has been proposed. Deferoxamine chelates iron in a catalytically inactive form, and thus inhibition by deferoxamine has been employed as evidence for the involvement of OH. generated by the Haber-Weiss reaction. We report here that deferoxamine also inhibits reactions catalyzed by the peroxidases of phagocytes, i.e., myeloperoxidase (MPO) and eosinophil peroxidase (EPO). The reactions inhibited include iodination in the presence and absence of chloride and the oxidation of guaiacol. Iodination by MPO and H2O2 is stimulated by chloride due to the intermediate formation of hypochlorous acid (HOCl). Iodination by reagent HOCl also is inhibited by deferoxamine with the associated consumption of HOCl. Iron saturation of deferoxamine significantly decreased but did not abolish its inhibitory effect on iodination by MPO + H2O2 or HOCl. Deferoxamine did not affect the absorption spectrum of MPO, suggesting that it does not react with or remove the heme iron. The conversion of MPO to Compound II by H2O2 was not seen when H2O2 was added to MPO in the presence of deferoxamine, suggesting either that deferoxamine inhibited the formation of Compound II by acting as an electron donor for MPO Compound I or that deferoxamine immediately reduced the Compound II formed. Iodination by stimulated neutrophils also was inhibited by deferoxamine, suggesting an effect on peroxidase-catalyzed reactions in intact cells. Thus deferoxamine has multiple effects on the formation and activity of phagocyte-derived oxidants and therefore its inhibitory effect on oxidant-dependent damage needs to be interpreted with caution.  相似文献   

6.
Myeloperoxidase catalyses the conversion of H2O2 and Cl- to hypochlorous acid (HOCl). It also reacts with O2- to form the oxy adduct (compound III). To determine how O2- affects the formation of HOCl, chlorination of monochlorodimedon by myeloperoxidase was investigated using xanthine oxidase and hypoxanthine as a source of O2- and H2O2. Myeloperoxidase was mostly converted to compound III, and H2O2 was essential for chlorination. At pH 5.4, superoxide dismutase (SOD) enhanced chlorination and prevented formation of compound III. However, at pH 7.8, SOD inhibited chlorination and promoted formation of the ferrous peroxide adduct (compound II) instead of compound III. We present spectral evidence for a direct reaction between compound III and H2O2 to form compound II, and for the reduction of compound II by O2- to regenerate native myeloperoxidase. These reactions enable compound III and compound II to participate in the chlorination reaction. Myeloperoxidase catalytically inhibited O2- -dependent reduction of Nitro Blue Tetrazolium. This inhibition is explained by myeloperoxidase undergoing a cycle of reactions with O2-, H2O2 and O2-, with compounds III and II as intermediates, i.e., by myeloperoxidase acting as a combined SOD/catalase enzyme. By preventing the accumulation of inactive compound II, O2- enhances the activity of myeloperoxidase. We propose that, under physiological conditions, this optimizes the production of HOCl and may potentiate oxidant damage by stimulated neutrophils.  相似文献   

7.
Myeloperoxidase induces apoptosis in src- or raxs-transformed fibroblasts, but not in parental nontransformed fibroblasts. This selectivity seems to be based on superoxide anion production by transformed cells, a recently described characteristic feature of transformed cells. Myeloperoxidase-mediated apoptosis induction is inhibited by SOD, catalase, 4-aminobenzoyl hydrazide, taurine and DMSO. This pattern of inhibition allows us to conclude that transformed cell derived superoxide anions dismutate to hydrogen peroxide, which fosters HOCl formation by myeloperoxidase. Hydrogen peroxide formation thereby is the rate-limiting step and depends on the cell density. In a second step, HOCl interacts with superoxide anions to yield the highly reactive apoptosis inducing hydroxyl radical. This conclusion was verified through selective apoptosis induction in transformed cells by direct addition of HOCl, which was also inhibited by SOD and DMSO. Our findings demonstrate a specific interplay between target cell derived superoxide anions and MPO during selective apoptosis induction.  相似文献   

8.
1,3-Butadiene was oxidized by human myeloperoxidase in the absence of KCl to yield butadiene monoxide (BM) and crotonaldehyde (CA), but at KCl concentrations higher than 50 mM, 1-chloro-2-hydroxy-3-butene (CHB) was the major metabolite detected; metabolite formation was dependent on incubation time, pH, KCl, 1,3-butadiene, and H2O2 concentrations. The data are best explained by 1,3-butadiene being oxidized by myeloperoxidase by two different mechanisms. First, oxygen transfer from the hemoprotein would occur to either C-1 or C-4 of 1,3-butadiene to form an intermediate which may cyclize to form BM or undergo a hydrogen shift to form 3-butenal, an unstable precursor of CA. Further evidence for this mechanism was provided by the inability to detect methyl vinyl ketone, a possible product of an oxygen transfer reaction to C-2 or C-3 of 1,3-butadiene, and by the finding that CA was not simply a decomposition product of BM under assay conditions. In the second mechanism, however, chloride ion is oxidized by myeloperoxidase to HOCl which reacts with 1,3-butadiene to yield CHB. Further evidence for this mechanism was provided by the finding that CHB was readily formed when 1,3-butadiene was added to the filtrate of a myeloperoxidase/H2O2/KCl incubation and when 1,3-butadiene was allowed to react with authentic HOCl. In addition, CHB was not detected when BM or CA was incubated with myeloperoxidase, H2O2, and KCl for up to 60 min, or when 1,3-butadiene and KCl were incubated with chloroperoxidase and H2O2 or with mouse liver microsomes and NADPH, enzyme systems which catalyze 1,3-butadiene oxidation to BM and CA, but unlike myeloperoxidase, do not catalyze chloride ion oxidation to HOCl. These results provide clear evidence for novel olefinic oxidation reactions by myeloperoxidase.  相似文献   

9.
Lactoperoxidase (LPO) is the major consumer of hydrogen peroxide (H(2)O(2)) in the airways through its ability to oxidize thiocyanate (SCN(-)) to produce hypothiocyanous acid, an antimicrobial agent. In nasal inflammatory diseases, such as cystic fibrosis, both LPO and myeloperoxidase (MPO), another mammalian peroxidase secreted by neutrophils, are known to co-localize. The aim of this study was to assess the interaction of LPO and hypochlorous acid (HOCl), the final product of MPO. Our rapid kinetic measurements revealed that HOCl binds rapidly and reversibly to LPO-Fe(III) to form the LPO-Fe(III)-OCl complex, which in turn decayed irreversibly to LPO Compound II through the formation of Compound I. The decay rate constant of Compound II decreased with increasing HOCl concentration with an inflection point at 100 μM HOCl, after which the decay rate increased. This point of inflection is the critical concentration of HOCl beyond which HOCl switches its role, from mediating destabilization of LPO Compound II to LPO heme destruction. Lactoperoxidase heme destruction was associated with protein aggregation, free iron release, and formation of a number of fluorescent heme degradation products. Similar results were obtained when LPO-Fe(II)-O(2), Compound III, was exposed to HOCl. Heme destruction can be partially or completely prevented in the presence of SCN(-). On the basis of the present results we concluded that a complex bi-directional relationship exists between LPO activity and HOCl levels at sites of inflammation; LPO serve as a catalytic sink for HOCl, while HOCl serves to modulate LPO catalytic activity, bioavailability, and function.  相似文献   

10.
The reaction between native myeloperoxidase and hydrogen peroxide, yielding Compound II, was investigated using the stopped-flow technique. The pH dependence of the apparent second-order rate constant showed the existence of a protonatable group on the enzyme with a pKa of 4.9. This group is ascribed to the distal histidine imidazole, which must be deprotonated to enable the reaction of Compound I with hydrogen peroxidase to take place. The rate constant for the formation of Compound II by hydrogen peroxide was 3.5.10(4) M-1.s-1. During the reaction of myeloperoxidase with H2O2, rapid reduction of added cytochrome c was observed. This reduction was inhibitable by superoxide dismutase, and this demonstrates that superoxide anion radicals are generated. When potassium ferrocyanide was used as an electron donor to generate Compound II from Compound I, the pH dependence of the apparent second-order rate constant indicated involvement of a group with a pKa of 4.5. However, with ferrocyanide as an electron donor, protonation of the group was necessary to enable the reaction to take place. The rate constant for the generation of Compound II by ferrocyanide was 1.6.10(7) M-1.s-1. We also investigated the reaction of Compound II with hydrogen peroxide, yielding Compound III. Formation of Compound III (k = 50 M-1.s-1) proceeded via two different pathways, one of which was inhibitable by tetranitromethane. We further investigated the stability of Compound II and Compound III as a function of pH, ionic strength and enzyme concentration. The half-life values of both Compound II and Compound III were independent of the enzyme concentration and ionic strength. The half-life value of Compound III was pH-dependent, showing a decreasing stability with increasing pH, whereas the stability of Compound II was independent of pH over the range 3-11.  相似文献   

11.
Oxidized low-density lipoprotein (LDL) is implicated in atherogenesis, and human atherosclerotic lesions contain LDL oxidized by myeloperoxidase, a heme protein secreted by activated phagocytes. Using hydrogen peroxide (H(2)O(2)), myeloperoxidase generates hypochlorous acid (HOCl), a powerful oxidant. We now demonstrate that HOCl produces sulfenamides, sulfinamides, and sulfonamides in model peptides, which suggests a potential mechanism for LDL oxidation and cross-linking. When we exposed the synthetic peptide PFKCG to HOCl, the peptide's thiol residue reacted rapidly, generating a near-quantitative yield of products. Tandem mass spectrometric analysis identified the products as the sulfenamide, sulfinamide, and sulfonamide, all formed by intramolecular cross-linking of the peptide's thiol and lysine residues. An intramolecular sulfinamide was also observed after the peptide PFRCG was exposed to HOCl, indicating that the guanidine group of arginine can also form a sulfur-nitrogen cross-link. The synthetic peptide PFVCG, which contains a free thiol residue but lacks nucleophilic amino acid side chains, formed an intermolecular sulfonamide when exposed to HOCl. Tandem mass spectrometric analysis of the dimer revealed that the free N-terminal amino group of one PFVCG molecule cross-linked with the thiol residue of another. This peptide also formed intermolecular sulfonamide cross-links with N(alpha)-acetyllysine after exposure to HOCl, demonstrating that the epsilon-amino group of a lysine residue can undergo a similar reaction. Moreover, human neutrophils used the myeloperoxidase-H(2)O(2) system to generate sulfinamides in model peptides containing lysine or arginine residues. Collectively, our observations raise the possibility that HOCl generated by myeloperoxidase contributes to intramolecular and intermolecular protein cross-linking in the artery wall. Myeloperoxidase might also use this mechanism to form sulfur-nitrogen cross-links in other inflammatory conditions.  相似文献   

12.
GSH is rapidly oxidized by HOCl (hypochlorous acid), which is produced physiologically by the neutrophil enzyme myeloperoxidase. It is converted into, mainly, oxidized glutathione. Glutathione sulfonamide is an additional product that is proposed to be covalently bonded between the cysteinyl thiol and amino group of the gamma-glutamyl residue of GSH. We have developed a sensitive liquid chromatography-tandem MS assay for the detection and quantification of glutathione sulfonamide as well as GSH and GSSG. The assay was used to determine whether glutathione sulfonamide is a major product of the reaction between GSH and HOCl, and whether it is formed by other two-electron oxidants. At sub-stoichiometric ratios of HOCl to GSH, glutathione sulfonamide accounted for up to 32% of the GSH that was oxidized. It was also formed when HOCl was generated by myeloperoxidase and its yield increased with the flux of oxidant. Of the other oxidants tested, only hypobromous acid and peroxynitrite produced substantial amounts of glutathione sulfonamide, but much less than with HOCl. Chloramines were able to generate detectable levels only when at a stoichiometric excess over GSH. We conclude that glutathione sulfonamide is sufficiently selective for HOCl to be useful as a biomarker for myeloperoxidase activity in biological systems. We have also identified a novel oxidation product of GSH with a molecular weight two mass units less than GSH, which we have consequently named dehydroglutathione. Dehydroglutathione represented a few percent of the total products and was formed with all of the oxidants except H2O2.  相似文献   

13.
We have studied the time course of the absorption of bovine liver catalase after pulse radiolysis with oxygen saturation in the presence and absence of superoxide dismutase. In the absence of superoxide dismutase, catalase produced Compound I and another species. The formation of Compound I is due to the reaction of ferric catalase with hydrogen peroxide, which is generated by the disproportionation of the superoxide anion (O-2). The kinetic difference spectrum showed that the other species was neither Compound I nor II. In the presence of superoxide dismutase, the formation of this species was found to be inhibited, whereas that of Compound I was little affected. This suggests that this species is formed by the reaction of ferric catalase with O-2 and is probably the oxy form of this enzyme (Compound III). The rate constant for the reaction of O-2 and ferric catalase increased with a decrease in pH (cf. 4.5 X 10(4) M-1 s-1 at pH 9 and 4.6 X 10(6) M-1 s-1 at pH 5.). The pH dependence of the rate constant can be explained by assuming that HO2 reacts with this enzyme more rapidly than O-2.  相似文献   

14.
Stimulation of the oxygen (O2) metabolism of isolated human neutrophilic leukocytes resulted in oxidation of hemoglobin of autologous erythrocytes without erythrocyte lysis. Hb oxidation could be accounted for by reduction of O2 to superoxide (O-2) by the neutrophils, dismutation of O-2 to yield hydrogen peroxide (H2O2), myeloperoxidase-catalyzed oxidation of chloride (Cl-) by H2O2 to yield hypochlorous acid (HOCl), the reaction of HOCl with endogenous ammonia (NH+4) to yield monochloramine ( NH2Cl ), and the oxidative attack of NH2Cl on erythrocytes. NH2Cl was detected when HOCl reacted with the NH+4 and other substances released into the medium by neutrophils. The amount of NH+4 released was sufficient to form the amount of NH2Cl required for the observed Hb oxidation. Oxidation was increased by adding myeloperoxidase or NH+4 to increase NH2Cl formation. Due to the volatility of NH2Cl , Hb was oxidized when neutrophils and erythrocytes were incubated separately in a closed container. Oxidation was decreased by adding catalase to eliminate H2O2, dithiothreitol to reduce HOCl and NH2Cl , or taurine to react with HOCl or NH2Cl to yield taurine monochloramine . NH2Cl was up to 50 times more effective than H2O2, HOCl, or taurine monochloramine as an oxidant for erythrocyte Hb, whereas HOCl was up to 10 times more effective than NH2Cl as a lytic agent. NH2Cl contributes to oxidation of erythrocyte components by stimulated neutrophils and may contribute to other forms of neutrophil oxidative cytotoxicity.  相似文献   

15.
Superoxide dismutase (SOD) enhanced the formation of hydroxyl radicals, which were detected by using the e.s.r. spin-trapping technique, in a reaction mixture containing 3-hydroxyanthranilic acid (or p-aminophenol), Fe3+ ions, EDTA and potassium phosphate buffer, pH 7.4. The hydroxyl-radical formation enhanced by SOD was inhibited by catalase and desferrioxamine, and stimulated by EDTA and diethylenetriaminepenta-acetic acid, suggesting that both hydrogen peroxide and iron ions participate in the reaction. The hydroxyl-radical formation enhanced by SOD may be considered to proceed via the following steps. First, 3-hydroxyanthranilic acid is spontaneously auto-oxidized in a process that requires molecular oxygen and yields superoxide anions and anthranilyl radicals. This reaction seems to be reversible. Secondly, the superoxide anions formed in the first step are dismuted by SOD to generate hydrogen peroxide and molecular oxygen, and hence the equilibrium in the first step is displaced in favour of the formation of superoxide anions. Thirdly, hydroxyl radicals are generated from hydrogen peroxide through the Fenton reaction. In this Fenton reaction Fe2+ ions are available since Fe3+ ions are readily reduced by 3-hydroxyanthranilic acid. The superoxide anions do not seem to participate in the reduction of Fe3+ ions, since superoxide anions are rapidly dismuted by SOD present in the reaction mixture.  相似文献   

16.
Production of superoxide ions by photosensitization of dyes   总被引:1,自引:0,他引:1  
The formation of superoxide anions by photosentization of fluorescein and other aromatic molecules has been observed using the enzyme lactoperoxidase as detector. Aerobic photosensitization caused the formation of Compound III with a rate of formation directly proportional to light intensity. Yields were such that photosensitization could be used as a superoxide anion generating process. Superoxide dismutase was also successfully used to demonstrate the specific involvement of superoxide ions in this photoprocess.  相似文献   

17.
Transformed fibroblasts are specifically eliminated by their nontransformed neighbors through intercellular induction of apoptosis. This process depends on the number of nontransformed effector cells and on the local density of transformed target cells. Intercellular signalling is inhibited by SOD (a scavenger of superoxide anions), taurine (a scavenger of HOCl), 4-aminobenzoyl hydrazide (a mechanism-based inhibitor of peroxidase), DMSO (a hydroxyl radical scavenger), and two inhibitors of NO synthase. Therefore, selective apoptosis induction seems to be based on superoxide anion production by transformed cells, their spontaneous dismutation to hydrogen peroxide, and HOCl generation by a novel effector cell-derived peroxidase. HOCl then interacts with target cell–derived superoxide anions to yield hydroxyl radicals. Due to the short diffusion pathway of superoxide anions, hydroxyl radical generation is confined to the intimate vicinity of transformed cells. In parallel, NO derived from effector cells interacts with superoxide anions of target cells to yield the apoptosis inducer peroxynitrite. Reconstitution experiments using transformed or nontransformed cells in conjunction with myeloperoxidase, HOCl, or an NO donor demonstrated that superoxide anions generated extracellularly by transformed cells participate in intercellular signalling and at the same time determine transformed cells as selective targets for intercellular induction of apoptosis.  相似文献   

18.
Myeloperoxidase uses hydrogen peroxide (H2O2) to generate hypochlorous acid (HOCl), a potent cytotoxic oxidant. We demonstrate that HOCl regulates the activity of matrix metalloproteinase-7 (MMP-7, matrilysin) in vitro, suggesting that this oxidant activates MMPs in the artery wall. Indeed, both MMP-7 and myeloperoxidase were colocalized to lipid-laden macrophages in human atherosclerotic lesions. A highly conserved domain called the cysteine switch has been proposed to regulate MMP activity. When we exposed a synthetic peptide that mimicked the cysteine switch to HOCl, HPLC analysis showed that the thiol residue reacted rapidly, generating a near-quantitative yield of products. Tandem mass spectrometric analysis identified the products as sulfinic acid, sulfonic acid, and a dimer containing a disulfide bridge. In contrast, the peptide reacted slowly with H2O2, and the only product was the disulfide. Moreover, HOCl markedly activated pro-MMP-7, an MMP expressed at high levels in lipid-laden macrophages in vivo. Tandem mass spectrometric analysis of trypsin digests revealed that the thiol residue of the enzyme's cysteine switch domain had been converted to sulfinic acid. Thiol oxidation was associated with autolytic cleavage of pro-MMP-7, strongly suggesting that oxygenation activates the latent enzyme. In contrast, H2O2 failed to oxidize the thiol residue of the protein or activate the enzyme. Thus, HOCl activates pro-MMP-7 by converting the thiol residue of the cysteine switch to sulfinic acid. This activation mechanism is distinct from the well-studied proteolytic cleavage of MMP pro-enzymes. Our observations raise the possibility that HOCl generated by myeloperoxidase contributes to MMP activation, and therefore to plaque rupture, in the artery wall. HOCl and other oxidants might regulate MMP activity by the same mechanism in a variety of inflammatory conditions.  相似文献   

19.
Recent studies have suggested that the inhibition of lymphocyte mitogenesis by D-penicillamine in the presence of copper could be mediated by the formation and action of hydrogen peroxide. To explore this possibility further, we first sought evidence of H2O2 generation by D-penicillamine in a cell-free system by a) measurement of copper-catalyzed D-penicillamine oxidation and the requirement for oxygen in this process; b) direct measurement of H2O2 formation during D-penicillamine oxidation by the peroxidase-mediated oxidation of fluorescent scopoletin; and c) evaluation of the possible synthesis of O2- during D-penicillamine oxidation. The addition of copper to D-penicillamine in physiologic buffer catalyzed D-penicillamine oxidation in a dose-dependent fashion. D-penicillamine oxidation was accompanied by O2 consumption with a molar ratio of approximately 2:1, but did not occur under anaerobic conditions. Furthermore, D-penicillamine oxidation resulted in the formation of amounts of H2O2 stoichiometrically equivalent to oxygen consumption (i.e., 1:1). Copper-catalyzed D-penicillamine oxidation caused reduction of nitroblue tetrazolium in a reaction blocked by superoxide dismutase, suggesting the formation of O2-. Additional studies confirmed that D-penicillamine inhibited PHA-induced mitogenesis of lymphocytes in the presence of copper, and that catalase protected the cells from this action. Furthermore, when polymorphonuclear leukocytes were incubated with D-penicillamine plus copper, hexose monophosphate shunt activity increased up to threefold with abrogation of this stimulation by catalase. None of the effects of D-penicillamine plus copper on cells were diminished by hydroxyl radical scavengers mannitol or benzoate. These results are consistent with oxygen-dependent copper-catalyzed oxidation of D-penicillamine in aqueous solutions leading to the formation of O2- and H2O2. H2O2 produced by this reaction can inhibit lymphocyte mitogenesis and stimulate neutrophil hexose monophosphate shunt activity in vitro and may be relevant to the therapeutic effects of D-penicillamine in vivo.  相似文献   

20.
Activated phagocytic cells generate hypochlorite (HOCl) via release of hydrogen peroxide and the enzyme myeloperoxidase. HOCl plays an important role in bacterial cell killing, but excessive or misplaced production of HOCl is also known to cause tissue damage. Studies have shown that low-molecular-weight thiols such as reduced glutathione (GSH), and sulfur-containing amino acids in proteins, are major targets for HOCl. Radicals have not generally been implicated as intermediates in thiol oxidation by HOCl, though there is considerable literature evidence for the involvement of radicals in the metal ion-, thermal- or UV light-catalysed decomposition of sulfenyl or sulfonyl chlorides which are postulated intermediates in thiol oxidation. In this study we show that thiyl radicals are generated on reaction of a number of low-molecular-weight thiols with HOCl. With sub-stoichiometric amounts of HOCl, relative to the thiol, thiyl radicals are the major species detected by EPR spin trapping. When the HOCl is present in excess over the thiol, additional radicals are detected with compounds which contain amine functions; these additional radicals are assigned to nitrogen-centered species. Evidence is presented for the involvement of sulfenyl chlorides (RSCl) in the formation of these radicals, and studies with an authentic sulfenyl chloride have demonstrated that this compound readily decomposes in thermal-, metal-ion- or light-catalysed reactions to give thiyl radicals. The formation of thiyl radicals on oxidation of thiols with HOCl appears to compete with non-radical reactions. The circumstances under which radical formation may be important are discussed.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号